A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com...

73
UNIVERSIDADE DA BEIRA INTERIOR Ciências A influência do processo de produção de scaffolds na regeneração óssea Mariana Costa da Silva Vallejo Dissertação para obtenção do Grau de Mestre em Biotecnologia (2º ciclo de estudos) Orientador: Prof. Doutor Ilídio Joaquim Sobreira Correia Covilhã, outubro de 2014

Transcript of A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com...

Page 1: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

UNIVERSIDADE DA BEIRA INTERIOR Ciências

A influência do processo de produção de

scaffolds na regeneração óssea

Mariana Costa da Silva Vallejo

Dissertação para obtenção do Grau de Mestre em

Biotecnologia

(2º ciclo de estudos)

Orientador: Prof. Doutor Ilídio Joaquim Sobreira Correia

Covilhã, outubro de 2014

Page 2: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

ii

I would like to dedicate my master thesis to my grandmother …

Page 3: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

iii

“The only place where success comes before work is in the dictionary”

Page 4: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

iv

Acknowledgements

First, I would like to thank my supervisor Professor Ilídio Correia for the opportunity to

develop my master's thesis with him. For all the dedication and all the time spent in carrying

out the whole project.

To Eng. Ana Paula from the Optics center of Universidade da Beira Interior for helping in the

acquisition of EDS and XRD results.

To Professor Abílio Silva for all the help and availability to perform the mechanical

characterization of the scaffolds.

To Professor Eugenia Gallardo, for providing several times the vacuum pump.

To Sónia Miguel, for helping with the acquisition of the SEM and CLSM images.

To all my group colleagues. One way or another all contributed to the development of my

work. A special thanks to Ricardo, Sofia and Elisabete, the people who spent more time with

me in the laboratory even in the after hours. For all the help and understanding throughout

this year.

To all my special friends for all their contribution. A special thanks to my boyfriend, David,

for all the patience and unconditional support through this year.

To the most important people in my life, my family, for all the love, motivation and support.

Mostly to my parents for all the sacrifices that they made this year to allow me to be here

today. You are all my strength.

Page 5: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

v

Abstract

The raising level of bone defects worldwide has become a major concern of public health.

Autografts, allografts and xenografts are some alternatives that are currently being used to

overcome bone related problems. However the risks of infection and immunological response

from the patient are very high. To surpass these handicaps, new biodegradable and

biocompatible structures (scaffolds) have been produced in the area of tissue engineering to

be used as temporarily bone substitutes. These structures allow cell adhesion and

proliferation, while giving mechanical support to the newly bone formation. Several

biocompatible materials have been used so far for scaffolds production. Nevertheless it is

fundamental that the right production method is chosen, since it will influence the

regenerative process. In this study a chitosan/alginate/β-TCP scaffold was produced by three

different techniques: rapid prototyping using a Fab@home 3D printer, freeze-drying and foam

replication method. The scaffolds were characterized by Fourier transformed Infrared

spectroscopy, X-ray diffraction, Energy dispersion Spectroscopy, Scanning Electronic

microscopy and water contact angle. Moreover, the porosity, mechanical properties, swelling

profile, and degradation behavior of the different scaffolds were also characterized. The

cytotoxic profile of the scaffolds was studied using a rezazurin assay. The cellular adhesion,

proliferation and internalization in the scaffolds were studied by optical, scanning electronic

and confocal laser scanning microscopy. The antimicrobial properties of the different

scaffolds were tested against Staphylococcus Aureus, using a Kirby-Bauer disk diffusion

method. All the results revealed that the chitosan/alginate/β-TCP scaffolds produced by

rapid prototyping had the most suitable properties for being applied in bone regeneration.

Keywords

Bone regeneration; Chitosan/Alginate/β-TCP scaffolds; Foam replication Method; Freeze-

drying; Rapid Prototyping

Page 6: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

vi

Resumo

O aumento de defeitos ósseos nas últimas décadas tem se tornado um problema de saúde a

nível mundial. Os autoenxertos, aloenxertos e xenoenxertos são algumas das alternativas

terapêuticas utilizadas para a reparação/regeneração dos defeitos do tecido ósseo. Contudo o

elevado risco de infeção e rejeição têm limitado a sua aplicação. A engenharia de tecidos tem

procurado soluções para ultrapassar estes problemas, nomeadamente atravéz da produção de

estruturas temporariamente substitutas do tecido ósseo que sejam biocompatíveis e

biodegradáveis. Estas estruturas são conhecidas como matrizes 3D (scaffolds), que permitem

a adesão e proliferação celular, fornecendo suporte mecânico até á formação de novo tecido

ósseo. O sucesso de um scaffold depende das suas propriedades químicas, mecânicas e

biológicas. Na produção destas estruturas tridimensionais têm sido usados vários materiais

biocompatíveis. Contudo também é importante ter em conta que a técnica usada na produção

dos scaffolds vai influenciar diretamente as suas propriedades e consequentemente a sua

ação na regeneração óssea. Neste estudo scaffolds de quitosano/alginato/β-TCP foram

produzidos usando três técnicas de produção diferentes: prototipagem rápida com uma

impressora 3D Fab@home, método de sublimação e método de replicação de esponja. Os

scaffolds produzidos foram caracterizados de forma a perceber qual a técnica que permitiria

a produção de scaffolds com propriedades adequadas para a regeneração óssea. Os diferentes

scaffolds foram caracterizados por espectroscopia de infravermelhos, difração de raio X,

espectroscopia de dispersão de energia, microscopia electrónica de varrimento e o carácter

hidrofílico foi caracterizado atravéz da determinação do ângulo de contacto. Por outro lado,

a porosidade e as propriedades mecânicas dos materiais foram analisadas, assim como o

inchaço e a sua degradação. O perfil citotóxico dos scaffolds foi avaliado usando um ensaio de

rezazurina e a adesão, proliferação e internalização celular nos scaffolds foi estudada por

microscopia ótica, microscopia electrónica de varrimento e microscopia confocal. As

propriedades antibacterianas foram avaliadas usando o método de difusão em placa de Kirby-

Bauer, usando a bactéria Staphylococcus Aureus como modelo. Os resultados obtidos

mostraram que os scaffolds de quitosano/alginato/β-TCP produzidos por prototipagem rápida

apresentaram propriedades adequadas para que futuramente possam ser usados na

regeneração óssea.

Page 7: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

vii

Palavras-chave

Método de replicação de esponja; Método de sublimação; Prototipagem rápida; Regeneração

óssea; Scaffolds de quitosano/alginato/β-TCP;

Page 8: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

viii

Resumo alargado

O osso humano é um tecido altamente resistente e vascularizado com capacidade de auto-

remodelação, e é responsável por várias funções no corpo humano. O osso é constituído pela

matriz orgânica (maioritariamente colagénio), matriz inorgânica (hidroxiapatite) e várias

células (osteoblastos, osteócitos e osteoclastos). No entanto, existem diferentes factores tais

como infeções, doenças ósseas e acidentes que conduzem ao aumento de defeitos ósseos, e

que nas últimas décadas se têm tornado um problema de saúde a nível mundial. Os

autoenxertos, aloenxertos e xenoenxertos são algumas das alternativas terapêuticas

utilizadas para a regeneração dos tecidos ósseos. No entanto, o elevado risco de infeção e

rejeição têm limitado a sua aplicação.

Neste contexto, a Engenharia de Tecidos surgiu como uma área promissora para a resolução

deste tipo de problemas. A Engenharia de Tecidos é uma área interdisciplinar que combina

biomateriais para desenvolver substitutos temporários biocompativeis de forma a restaurar,

manter e melhorar a função dos tecidos biológicos. Um dos conceitos da engenharia de

tecidos passa por produzir matrizes 3D (scaffolds) que permitem a adesão e proliferação

celular e fornecem suporte mecânico até à formação de novo tecido ósseo. O sucesso de um

scaffold depende de diversos parâmetros tais como biocompatibilidade, biodegradabilidade,

hidrofobicidade, porosidade e resistência mecânica, com o objetivo de promover o

crescimento, proliferação e intenalização celular. Na produção de scaffolds têm sido usados

vários materiais para a regeneração óssea. Estes materiais incluem cerâmicas e polímeros

(naturais e sintéticos) que conferem resistência e biocompatibilidade aos scaffolds. Contudo,

também é importante ter em conta a técnica usada na produção dos scaffolds. O processo de

produção pode influenciar diretamente as propriedades fisico-químicas, mecânicas e

biológicas de um scaffold e, consequentemente, a sua ação na regeneração óssea. Desta

forma, é essencial encontrar uma técnica que permita a produção de scaffolds com as

melhores propriedades para a regeneração óssea.

Neste estudo scaffolds de quitosano/alginato/β-TCP foram produzidos usando três técnicas de

diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de

replicação de esponja. Após a produção dos três tipos de scaffolds, estes foram

caracterizados de forma a perceber qual a técnica que permitiria a produção de scaffolds

com propriedades adequadas para a regeneração óssea. A análise macroscópica dos scaffolds

produzidos revelou que a técnica de prototipagem rápida foi a única que permitiu o controlo

da estrutura e dimensões do scaffold. A estrutura interna e proriedades físico-químicas dos

scaffolds foram caraterizados por microscopia eletrónica de varrimento, espectroscopia de

Infravermelho (FTIR), difração de raios-X (XRD) e ainda foram determinados os ângulos de

contato na superfície dos scaffolds. Os resultados obtidos de XRD permitiram concluir que

Page 9: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

ix

apesar dos diferentes processos de produção, os três tipos de scaffolds não sofrem alterações

na sua estrutura cristalina. No entanto, os resultados obtidos por FTIR permitiram concluir

que os scaffolds produzidos pelo método de replicção de esponja sofreram a perda do seu

conteúdo polimérico. Todos os scaffolds produzidos apresentaram um carater hidrofílico, o

que é fundamental para promover adesão e proliferação celular.

Os resultados da caracterização mecânica dos scaffolds revelaram que os scaffolds

apresentaram valores de resistência à compressão e módulos de Young mais próximos

daqueles que o osso trabecular humano possui. Para além disso, todos os scaffolds

apresentaram tamanhos de poro desejáveis para a internalização e proliferação celular.

O perfil citotóxico dos scaffolds foi avaliado através de um ensaio de rezazurina, usando

osteoblastos humanos. Nenhum dos scaffold apresentou citotoxicidade. A adesão e

internalização celular nos diferentes scaffolds foram analisadas por microscopias ótica,

electrónica de varrimento e confocal. Os resultados revelaram que os osteoblastos humanos

aderiram e proliferaram na superfície dos scaffolds, assim como migraram e proliferaram no

interior dos biomaterias produzidos.

As propriedades antibacterianas dos diferentes scaffolds foram testadas através do teste de

difusão em placa de Kirby-Bauer, usando como modelo a bactéria gram-positiva

Staphylococcus Aureus. Os resultados obtidos revelaram que os scaffolds produzidos por

prototipagem rápida e método de sublimação possuem propriedades antimicrobianas.

Com base nos resultados obtidos concluiu-se que os scaffolds de quitosano/alginato/β-TCP

produzidos pela técnica de prototipagem rápida foram os que apresentaram melhores

propriedades para que futuramente possam serem aplicados na regeneração do tecido ósseo.

Page 10: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

x

Table of contents

CHAPTER I - INTRODUCTION ......................................................................... 1

1.1 Bones in the human body ................................................................................ 2

1.1.1 Types of bones ....................................................................................... 2

1.1.2 Bone matrix .......................................................................................... 3

1.1.3 Bone cells ............................................................................................. 4

1.1.4 Bone remodeling .................................................................................... 5

1.2 Bone disorders ............................................................................................. 6

1.3 Bone grafts ................................................................................................. 7

1.4 Tissue engineering ........................................................................................ 7

1.4.1 Biomaterials used for bone scaffolds production .............................................. 8

1.4.2 Scaffolds properties............................................................................... 11

1.4.3 Techniques used for the production of scaffolds ............................................ 12

1.5 Aims ....................................................................................................... 14

CHAPTER II - MATERIALS AND METHODS ..................................................... 15

2.1 Materials .................................................................................................. 16

2.2 Methods ................................................................................................... 16

2.2.1 Production of the scaffolds ...................................................................... 16

2.2.1.1 Production of the scaffolds by the Foam Replication Method (FRM) .................. 16

2.2.1.2 Production of the scaffolds by the Freeze-Drying technique ........................... 17

2.2.1.3 Production of the scaffolds by Rapid Prototyping ........................................ 17

2.2.2 Morphological and physicochemical characterization of the scaffolds .................. 17

2.2.2.1 Morphological analysis of the scaffolds ..................................................... 17

2.2.2.2 Fourier Transform Infrared Spectroscopy analysis ........................................ 17

2.2.2.3 X-Ray diffraction analysis ..................................................................... 18

2.2.2.4 Energy dispersive spectroscopy .............................................................. 18

2.2.3 Water contact angle measurement ............................................................ 18

Page 11: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

xi

2.2.4 Total porosity evaluation ........................................................................ 18

2.2.5 Characterization of the swelling capacity of the scaffolds ................................ 19

2.2.6 Characterization of the degradation rate of the scaffolds ................................ 19

2.2.7 Mechanical characterization of the different scaffolds .................................... 20

2.2.8 Biological characterization of the chitosan/alginate/β-TCP scaffolds .................. 20

2.2.8.1 Characterization of the cytotoxic profile of the scaffolds .............................. 20

2.2.8.2 Proliferation of the osteoblast cells in the presence of the scaffolds ................ 21

2.2.8.3 Confocal Laser Scanning Microscopy ........................................................ 22

2.2.9 Evaluation of the antibacterial activity of the scaffolds ................................... 22

2.2.10 Statistical Analysis ............................................................................... 22

CHAPTER III - RESULTS AND DISCUSSION ..................................................... 23

3.1 Macroscopic properties of the scaffolds ............................................................ 24

3.2 Physicochemical characterization of the scaffolds ............................................... 27

3.2.1 Fourier Transform Infrared Spectroscopy ..................................................... 27

3.2.2 X-Ray diffraction analysis ........................................................................ 28

3.2.3 Energy Dispersive Spectroscopy ................................................................ 29

3.3 Porosity of the scaffolds ............................................................................... 30

3.4 Water contact angle determination ................................................................. 31

3.5 Swelling capacity of the scaffolds ................................................................... 32

3.6 Degradation behavior of the scaffolds .............................................................. 33

3.7 Mechanical characterization of the scaffolds ...................................................... 35

3.8 Biological characterization of the scaffolds ........................................................ 37

3.8.1 Evaluation of the cytotoxic profile of the scaffolds ........................................ 37

3.8.2 SEM and CLSM images of the cells in contact with scaffolds .............................. 40

3.9 Antibacterial activity of the scaffolds ............................................................... 42

3.10 Summary ................................................................................................ 42

CHAPTER IV - CONCLUSIONS AND FUTURE PERSPECTIVES ........................... 44

Page 12: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

xii

CHAPTER V - BIBLIOGRAPHY ....................................................................... 46

Page 13: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

xiii

List of figures

Figure 1: Different types of bones found in the human body. ........................................ 2

Figure 2: Representation of the composition of the bone matrix .................................... 3

Figure 3: Schematic representation of the main phases of bone remodeling ...................... 5

Figure 4: Schematic representation of the role of a scaffold in bone regeneration. ............. 8

Figure 5: Ilustration of the chemical structure of alginate ............................................ 9

Figure 6: Ilustration of the chemical structure of chitosan ......................................... 10

Figure 7: Ilustration of the chemical structure of HA and β-TCP ceramics. ...................... 11

Figure 8: Macroscopic images of the produced scaffolds ............................................ 25

Figure 9: SEM images of the internal structure of the scaffolds .................................... 26

Figure 10: FTIR analysis of the 3D scaffolds ............................................................ 27

Figure 11: X-Ray spectra of β-TCP and 3D scaffolds .................................................. 29

Figure 12: Total porosity the different scaffolds ...................................................... 31

Figure 13: Swelling behavior of the different scaffolds. ............................................. 33

Figure 14: Degradation rate of the different scaffolds............................................... 34

Figure 15: Characterization of the scaffolds compressive strength ................................ 36

Figure 16: Characterization of the scaffolds Young’s Modulus ..................................... 36

Figure 17: Evaluation of the cytotoxic profile of the scaffolds ..................................... 37

Figure 18: Macroscopic images of the cells in the presence of the scaffolds .................... 39

Figure 19: SEM images of the cells at the scaffolds surface ......................................... 40

Figure 20: CLSM images of the cells distribution within the scaffolds structure ............... 41

Page 14: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

xiv

Figure 21: Evaluation of the scaffolds antibacterial activity trough the Kirby-Bauer diffusion

method .................................................................................................. 42

Page 15: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

xv

List of tabels

Table 1: Young´s modulus and compressive strength values of human cortical and trabecular

bone ....................................................................................................... 4

Table 2: EDS analysis of the produced 3D scaffolds ................................................... 30

Table 3: Water contact angle values determined for the different produced scaffolds........ 32

Table 4: Summary of the properties of the chitosan/alginate/β-TCP scaffolds ................. 43

Page 16: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

xvi

Acronyms

3D Three-Dimensional

Ba2+ Barium

β-TCP βeta-Tricalcium Phosphate

BMP Bone Morphogenetic Proteins

BSA Bovine Serum Albumin

Ca2+ Calcium

CAD Computer-Aided Design

Cbfal Core binding factor

CLSM Confocal Laser Scanning Microscopy

dethanol Ethanol´s density

DMEM-F12 Dulbecco’s Modified Eagle’s Medium

ECM Extracellular Matrix

EDS Energy Dispersive Spectroscopy

EDTA Ethylenediaminetetraacetic Acid

EtOH Ethanol

FBS Fetal Bovine Serum

FTIR Fourier-Transform Infrared Spectroscopy

FRM Foam replication method

HA Hydroxyapatite

hOB Human osteoblasts

MSC Mesenchymal Stem Cells

OB Osteoblasts

OC Osteoclasts

OT Osteocytes

PBS Phosphate-Buffered Saline

PCL Poly (ε-caprolactone)

PLGA Poly (lactic-co-glycolic acid)

PLLA Poli (L-lactic-acid)

PI Propidium Iodide

PTH Paratiroid Hormone

PVA Poly(vinyl) Alcohol

RANKL NF-kappa B ligand

RP Rapid prototyping

RT Room Temperature

S.Aureus Staphylococcus Aureus

SEM Scanning Electron Microscopy

Page 17: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

xvii

SFF Solid free form fabrication

SLS Selective Laser Sintering

Sr2+ Strontium

STL Stereolithography

TE Tissue Engineering

TNF Tumor Necrosis Factor

Ts Resistance to Compression

XRD X-Ray Diffraction

YM Young’s Modulus

Page 18: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

Chapter I

Introduction

Page 19: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

2

1. Introduction

1.1 Bones in the human body

The skeleton is one of the most important organs of the human body. By adult age, it is

composed by 206 bones, and it is involved in the body sustention and locomotion [1].

Furthermore, it is responsible by other body functions, including the protection of vital organs

(brain, lungs, and heart), and is also involved in the production of blood cells through

hematopoiesis, as well as the storage of important ions, such as phosphorus and calcium [2].

1.1.1 Types of bones

Bones can present various shapes (Figure 1a). The long bones, like the femur and tibia, are

rigid and dense bones, providing strength, structure and mobility, supporting most of the

human weight [3]. On the other hand, short bones like the carpal bones of the wrist, or the

tarsal bones of the ankle, are responsible for support and stability in small movements [1, 4].

Flat bones, like the skull, present a flat and broad surface, important for muscle attachment,

as well as organ protection [1, 2]. Finally, irregular bones, like the vertebrae, are named due

to their variety of shapes and present many surface features for muscle or articulation

attachment [1].

Figure 1: Different types of bones found in the human body (a). Distribution of cancellous and compact

bone tissue on flat bones (b) (adapted from [1]).

Page 20: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

3

1.1.2 Bone matrix

Bone tissue is composed by bone matrix and cells. Bone matrix is formed by organic and

inorganic phases (figure 2), in a ratio of 35/65% [1]. The inorganic phase is mainly composed

of hydroxyapatite (HA), while the organic phase consists primarily of collagen and

proteoglycans. These two components are responsible for the majority of the functional

characteristics of bone: while collagen confers elasticity to the matrix, HA provides

compression strength [1, 3, 5]. In a healthy bone, both the inorganic and organic components

must exist in a fine tuned balance: a decrease in collagen composition leads to excessive

mineralization, making the bone brittle, while a decrease in the mineral content can make

the bone excessively flexible [1].

Figure 2: Representation of the composition of the bone matrix at different scales, from macro to nano

perspective (adapted from [6]).

In relation to the density of the bone matrix, bones can be classified as cortical/compact or

trabecular/cancellous (figure 1b) [1, 4]. Cortical bone is a dense and hard tissue and accounts

for 80% of the skeleton, with approximately 10% porosity, and high compressive strength. On

the other hand, trabecular bone presents a higher porosity, around 90%, with lower

mechanical resistance, representing near 20% of the skeleton [5, 7, 8]. The mechanical

characteristics of these types of bone tissue are represented in table 1.

Page 21: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

4

Table 1:Young´s modulus and compressive strength values of human cortical and trabecular bone

(adapted from [7]).

Young Modulus (MPa) Compressive Strengh (MPa)

Cortical bone

10000 - 25000

100 - 200

Trabecular bone

60 - 3200

2 - 18

1.1.3 Bone cells

Bone tissue is surrounded and populated by a variety of different cells. Bone tissue

functionality is maintained by three main cells types: osteoblasts (OB), osteocytes (OT) and

osteoclasts (OC) [9].

OB are basophilic, mononuclear cells, derived from mesenchymal steam cells (MSC), and are

present at the surface of the bone [9, 10]. These cells play an important role in the formation

of bone matrix (figure 3b), being responsible for the synthesis and secretion of collagen type I

and proteoglycans [10]. Furthermore, OB produce vesicles containing enzymes and important

ions (phosphate and calcium) that are released by exocytose, and are responsible for the

formation of hydroxyapatite crystals [9, 10]. After the deposition of the organic matrix, the

OB lay down a pre-mineralized matrix, called osteoid.

OT are mature, long-lived cells and the most abundant cell type in the bone matrix (85-90%).

These cells possess cytoplasmatic prolongations (filopodia), creating a complex intercellular

network (figure 3c) and maintaining its integrity [1, 11]. OT are also responsible for

intercellular communication, forming a complex cellular network, which allows the exchange

of nutrients and oxygen between the blood vessels and distant osteocytes [11]. Moreover,

these cells are also responsible for osteocytic osteolysis, breaking down the bone matrix, and

releasing calcium that is essential for calcium homeostasis [12].

OC are multinuclear and polymorphic giant cells that derive from hematopoietic stem cells

(figure 3a) [13]. They are involved in the matrix reabsorption and are responsible for the bone

demineralization, by promoting an increase in the solubility of HA crystals, as well as being

involved in the enzymatic destruction of the organic matrix [10]. Furthermore, these cells are

rich in actin and myosin, which are crucial for OC attachment to the bone surface and then by

bone resorption [9].

Page 22: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

5

1.1.4 Bone remodeling

One of the bone´s most important characteristics is its self-regeneration capability. This

tissue is in constant renovation, being the old bone reabsorbed and new one formed. Bone

remodeling comprises three main phases: resorption, reversal and formation (figure 3) [14].

Figure 3: Representation of the main phases of bone remodeling (adapted from [9]).

1.1.4.1 Resorption Phase

Bone resorption involves the degradation of the inorganic and organic components, through

different processes [4, 5, 14]. An increase in the parathyroid hormone (PTH) levels initiates

the bone resorption process, promoting the surface expression of the receptor of nuclear

factor-kappa B ligand (RANKL) by osteoblasts. At the same time, the osteoclastogenic

stimulus induces the expression of nuclear factor-kappa B (RANK) by pre-osteoclastic cells.

This RANK/RANKL interaction promotes the migration of OC precursors to the bone surface,

and subsequent maturation and differentiation into multinucleated OC [5, 10].

Simultaneously, PTH stimulates OB to produce proteins, such as collagenase, that are

involved in the degradation of collagen present in the bone matrix [1]. Finally, the

differentiated OC attach to the bone surface [11], and using specific proton and chloride

pumps (ATP-consuming vacuolar proton pumps) [15], they release hydrochloric acid that

Page 23: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

6

dissolves HA crystals, allowing proteolytic enzymes to reach and degrade the collagen in the

bone matrix [5].

It is important to consider that a continuous increase in the expression of RANKL promotes a

decrease in the osteoprogesterin (OPG) levels [1, 9]. OPG is a competitive inhibitor of RANK

that also binds to the RANKL. With an increase in PTH levels, the RANKL levels increase and

OPG levels decrease, enhancing RANK/RANKL interaction, OC migration and differentiation,

and promoting bone resorption.

1.1.4.2 Reversal and formation phases

After the resorption phase, bone enters a reversal state, with OC leaving the bone surface

and suffering apoptosis [5, 14]. Collagen remnants at the bone surface are removed by

mononuclear cell of undetermined lineage, preparing the bone surface for subsequent

osteoblast-mediated bone formation [16]. This begins the process of bone formation [14]. OB

precursors are attracted to the reabsorbed site and suffer differentiation into OB (figure 3 b).

After OB recruitment, they start to produce and lay down the organic matrix, called osteoid.

Posteriorly, some OB get trapped inside the new matrix and suffer differentiation into OT.

These cells will be part of the bone matrix structure, promoting its maintenance (figure 3 c)

[5, 14]. Finally, in the last osteoid phase, bone formation stops and following bone

mineralization, bone lining cells remain in a quiescent state, as it is illustrated in Figure 3d.

The remodeling cycle is fundamental for the maintenance of the integrity of the bone. For

that, an equilibrium between the matrix formation and resorption is necessary.

1.2 Bone disorders

Apart from its high resistance and self-regeneration capability, bone can suffer anomalous

remodeling processes or injuries that can be caused by different factors, such as trauma,

infections and tumors. Age, smoking and diabetes can also hinder bone remodeling and

regeneration [4, 17-19].

Osteoporosis is the bone remodeling disease with the highest prevalence, generally affecting

females, and progresses with age [17]. According to the World Health Organization (WHO),

osteoporosis is characterized by low bone mineral density, which causes bone fragility,

reduced strength and increased vulnerability to fractures [18]. Moreover, the number of hip

fractures worldwide due to osteoporosis is expected to rise three-fold by the middle of the

next century, from 1.7 million in 1990 to 6.3 million by 2050 [20].

Page 24: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

7

Another common bone pathology is Paget’s disease, being the second most common disorder

of bone remodeling, after osteoporosis. Its impact usually increases with age, and it affects

men more frequently. This disease is characterized by heightened presence of OC, resulting

in dramatic increases in bone reabsorption in the affected bones and, therefore, decreasing

the bone strength [18, 19].

Despite these diseases, other bone fractures can occur due to pedestrian and transport

accidents or even assaults [21].

1.3 Bone grafts

Nowadays different therapeutic strategies have been studied, in order to overcome bone

damages and improve treatment options. The most common therapies for the treatment of

large bone defects are bone grafts. Depending on their source, they can be classified as

autologous, allogenic or xenogenic grafts [22, 23].

Autologous grafts are obtained from the patient’s own body. These grafts have the advantage

of presenting low rejection risk. However, they can cause additional pain, and there is the

risk of donor site infection. In addition, autologous bone grafts have limited supply, requiring

the use of allogeneic grafts for extensive damage [24]. Allogeneic grafts are isolated from

donors of the same species, leading to a higher risk of immune rejection when compared to

the autologous grafts [22, 24]. Another approach used so far is xenogeneic bone grafting,

which is obtained from animals from different species. Such grafts can cause infections,

immune rejection and present limited functionality. Another option is to use decellularized

matrix, providing a suitable support for cellular repopulation, while eliminating problems

associated with xenogeneic cells [23, 24].

In order to overcome the drawbacks associated with the use of bone grafts, several bone

substitute materials have been studied so far, based on the concept of tissue engineering [6,

22].

1.4 Tissue engineering

Tissue engineering (TE) is an interdisciplinary field of research that applies the principles of

engineering and life sciences, to develop biological substitutes to restore, maintain, or

improve tissues functions [8, 24-27]. Different TE approaches have been used to create new

therapeutics for the regeneration of bone defects including drug delivery systems and

scaffolds, that are aimed to promote the regeneration of bone without causing adverse

Page 25: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

8

effects on the patient [28]. Scaffolds can be defined as material-based three-dimensional

structures, that intent to improve bone healing, while they are reabsorbed by the body

(Figure 4) [23, 27, 29].

Figure 4: Schematic representation of the role of a scaffold in bone regeneration (adapted from [30]).

1.4.1 Biomaterials used for bone scaffolds production

Metals, polymers and ceramics are the most widely used materials for the production of bone

scaffolds [6, 22, 31, 32]. Metals, like titanium and aluminum, have been used for their high

mechanical strength, light-weight nature and biocompatibility [33-35], being used for a wide

variety of applications, such as screws and fixation devices [35]. However, they can release

metallic ions and suffer corrosion, causing some toxicity in the body, as well as being poorly

osteointegrated in the surrounding bone [32, 35].

Polymers can be divided in synthetic and natural [35, 36]. Synthetic polymers have been used

in bone TE due to their biocompatibility, biodegradability, and easiest manufacturing into

different shapes. These polymers include polyurethane (PU), polypropylene (PP) and

polyesters like polylactic acid (PLA), polyglycolic acid (PGA) and poly (lactic-co-glycolic) acid

(PLGA) [37]. However, these materials also present disadvantages, such as release of by-

products that can lead to some local temporary disturbances, or may result in inflammatory

reactions [38]. Natural polymers have also been used in bone TE due to their several

advantages, including good biocompatibility, biodegradability and enhanced cell adhesion and

Page 26: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

9

proliferation. Among the most used natural polymers are included collagen, silk fibroin,

hyaluronic acid, cellulose, dextran, chitosan and alginate [37].

Alginate is a well-known polysaccharide that has been studied over the years [28, 39-42]. This

natural anionic polymer exhibits some of the desirable properties for bone TE, including good

biocompatibility, biodegradation, gel-forming capacity (through the addition of divalent

cations, like Ca2+, Ba2+ and Sr2+) and low cost [43]. Alginate is obtained from brown sea

weeds and is composed by 1,4-linked β-D-mannuronic acid (M) and α-L-guluronic acid (G)

residues (figure 5), which allow the polymer to create stable hydrogels through ionic

interactions between its functional groups [28]. However, the properties of the alginate gels

(stiffness, elasticity and stability) are highly sensitive to the species and concentrations of

ions present in solution. When the content of the G residues is higher than the M residues,

brittle gels are obtained [43, 44]. Furthermore, cells do not bind directly to the negatively

charged polymer [41, 43].

Figure 5: Chemical structure of alginate and organization of the M and G blocks in its structure.

(adapted from [42]).

In order to produce more stable and consistent gels, and to improve cells adhesion to

alginates, it has been ionically associated with cationic polymers, namely chitosan [41, 43,

45].

Chitosan is a natural polysaccharide composed by N- acetetylglucosamine and N-glucosamine

groups randomly distributed through its biopolymer chain (figure 6) [46]. This polymer is

produced through the deacetylation of marine chitin [47], being available with different

degrees of deacetylation (percentage of free amine groups), viscosity and molecular weight

[48]. Chitosan is only soluble in acidic solutions, due to the protonation of its amine groups

[49]. Moreover, these positively charged groups interact with the negatively charged

phospholipids of cells membranes, promoting cellular adhesion, proliferation and

differentiation [50]. In addition, chitosan interaction with the electronegative residues

present at the bacteria surface, leads to decrease of the bacterial cell wall permeability, and

thus bacteria death [47, 49]. Furthermore, due to its biocompatibility and biodegradability

Page 27: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

10

properties, low cost and availability, chitosan has attracted great attention in bone TE [46,

47, 49, 51-53].

Figure 6: Chemical structure of (A) chitosan and (B) protonated chitosan (adapted from [51]).

Different studies have reported the production of scaffolds containing both chitosan and

alginate, for bone tissue engineering [39, 41, 43, 52-56]. However, these scaffolds do not

present the desired mechanical properties for supporting load bearing applications [6, 29,

56]. The combination of polymers with ceramics can improve the mechanical properties of

the scaffolds [41, 56-59].

Ceramics are known for having high compressive strength, good bioresorbability and

biocompatibility [60]. The ceramics currently used for preparing scaffolds for bone

regeneration are calcium phosphate (CaP) based bioceramics. Among these, hydroxyapatite

(HA) and beta-tricalcium phosphate (β-TCP) are the most commonly used in clinical

applications [61]. Since HA (figure 7) is the major component of bone, it is usually the first

choice for scaffolds production [62]. There are several studies where this ceramic was

combined with chitosan and alginate to produce bone scaffolds [41, 56, 57, 63]. However,

this ceramic presents some disadvantages like high cost and low biodegradability [64].

Conversely, Β-TCP presents better biodegradability, osteocondutivity and osteoindutivity.

Moreover, β-TCP is cheaper than HA [32, 63, 64] and has been used in several studies for bone

regeneration [58, 63, 65-67]. However, β-TCP also presents some limitations like brittleness

and poor resistance to fatigue [63].

Page 28: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

11

Figure 7: Representation of the different HA (A) and β-TCP (B) compositions.

There are several studies revealing that the combination of both bioactive ceramics and

natural polymers improves the mechanical and biological properties of the scaffolds [41, 56-

59]. In fact, Qiao et al developed an injectable paste containing chitosan, alginate and

calcium phosphate for bone regeneration [39], demonstrating that the combination of these

components is very promising for bone TE applications.

1.4.2 Scaffolds properties

An ideal scaffold to be applied in bone TE has to meet certain parameters [8, 27, 29, 36, 38,

68-70] , such as:

Biocompatibility and biodegradability: The biocompatibility of a scaffold can be defined as

the ability to perform its function in the host tissue without eliciting any immune response

[36, 38]. This is very important, because it allows the scaffold to serve as a permanent or

temporary structure to support the formation of new tissue [36]. Biodegradability is defined

as the tunable rate of degradation, matching the growth of new bone tissue, replacing the

scaffold by new bone tissue. Such will allow the transference of the mechanical load to the

new tissue [27, 36].

Porosity and pore size: Porosity is an essential property in a scaffold aimed for bone

regeneration. An interconnecting porous structure increases the surface area of the scaffold,

allowing increased cell growth at the scaffold surface and within its structure [69]. It also

maximizes cell-to-cell interactions inside the scaffold. In addition, pores also enable the

diffusion of nutrients and gases to the entrapped cells [27]. According to Karageorgiou et al,

the minimum pore size required for bone regeneration is 100 µm. However, the increase of

the material’s porosity may lead to the reduction of the mechanical properties of the

Page 29: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

12

scaffolds [69]. It is also important to consider the type of bone that the scaffold is aimed to

replace, since the trabecular and cortical bones present different ranges of porosity [1].

Mechanical properties: The mechanical properties of human bones vary according to the

type of bone (table 1). Cortical bone has a compressive strength of 100-230 MPa and

trabecular bone has a compressive strength of 2-18 MPa. The young´s modulus of cortical and

trabecular bones vary between 10000-25000 MPa and 60-3200 MPa, respectively [7].

Therefore, a scaffold for human bone regeneration must have suitable mechanical properties

to support the mechanical loads on the damaged area, while maintaining the biological

properties required for cell growth and new matrix formation [29].

Surface properties: The scaffolds surface properties include the hydrophobic/hydrophilic

character, charge, roughness/softness and chemical composition of the materials. Such

properties play an essential role on cellular adhesion, morphology, differentiation and

proliferation [68].

1.4.3 Techniques for the production of scaffolds

The selection of the proper materials for the production of bone scaffolds is essential.

However, the choice of the right manufacturing technique is also of great importance [70].

Since there are different production parameters (like pressure, temperature and time)

influencing the properties of the scaffolds, the fabrication processes play an important role in

the final characteristics of the biomaterials [31, 70]. Therefore, it is essential to understand

and control the processes variations [32, 70]. Different production methods have been used

through the years for the production of scaffolds for bone regeneration, including Foam

Replication Method, Freeze-drying and Rapid Prototyping [12, 36, 41, 60, 63, 70-72].

In Foam Replication Method (FRM) a desired solution is impregnated in a polymeric template

[63, 73-75]. The exposure to high temperatures, known as sintering, promotes the template

melting and the fusion of the solution particles [63, 76-78], resulting in a final 3D scaffold

with a structure similar to the initial template [64, 76, 77]. The strength of the construct is

critically dependent on the ability to achieve a solid network with high density, thus

depending on the particle packing of the infiltrated foam [75]. Previous studies reported the

use of this technique in the production of scaffolds composed by glasses [73, 75] and ceramics

[63, 74]. Torres et al describes this technique as a simple and cost-effective method that

allows good reproducibility [63]. Furthermore, using a foam with an appropriate architecture,

scaffolds with a microstructure similar to trabecular bone can be prepared by FRM [63, 75].

However, the scaffold final shape is restricted to the initial template structure [63], and

sintering temperature must be optimized according to the desired materials [60]. For

Page 30: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

13

instance, it is reported that the sintering temperature of scaffolds containing β-TCP has to be

below 1200ºC, or else it will acquire its α-TCP form [60].

Another method usually applied in the production of bone scaffolds is Freeze-drying [26, 41,

72, 79]. This technique is based on the freezing of a desired solution, followed by a process of

sublimation. This process allows the removal of solvents, under the action of vacuum,

resulting in porous scaffolds [26, 80]. By varying the cooling rate, phase separation could

result in different porosities of the scaffold [81]. Several studies using this technique have

been reported for the production of chitosan [82, 83], chitosan/alginate [41, 84], and

chitosan/alginate/HA scaffolds [41]. This production technique is known for being cheap and

versatile [26, 72, 79]. Furthermore, it is possible to produce scaffolds with high porosity,

which is crucial for bone TE applications. The disadvantages of this technique are associated

with the Iack of control in the structure and shape of the final scaffold [8].

More recently, techniques based on rapid prototyping (RP) have been attracted a lot of

attention [31, 38, 71, 80, 85]. In RP techniques, 3D models are designed by computer-aided

design (CAD), and created through a layer by layer process until the desired 3D structure is

completed [58, 86, 87]. This technique allows the production of different structures for

several applications (bone, cartilage, heart valves, nerves) in TE, in short periods of time.

Furthermore, it is reported that this technique promotes the desired level of complexity into

the scaffold that cannot be achieved with the conventional techniques [35, 36, 71].

Fab@home 3D printing system is an example of RP that has the ability of producing structures

composed by different materials like plastics and hydrogels. Previous studies using this

technique, reported the production of alginate [88] and alginate/β-TCP scaffolds [58, 65] for

bone TE. According to Diogo et al, RP is relatively simple and inexpensive technique.

Furthermore, the Fab@home model has advantages over other equipments, since it allows the

employment of different materials, such as composite or viscous solutions [89]. However, the

production of scaffolds using this technique depends on the optimization of material viscosity

and printing parameters [58].

One of the challenges in bone TE is to find a technique that allows the production of scaffolds

with suitable characteristics for being applied in bone tissue regeneration [25, 80]. There are

several reviews that report the production of bone scaffolds, using different materials,

through distinct techniques [25, 27, 38, 80]. In fact, these reviews include the advantages

and disadvantages of several techniques in the production of scaffolds for bone TE.

Nevertheless, it is important to understand how the different production processes can

influence the properties of the same scaffold and their impact in bone regeneration.

Therefore, the main purpose of the present study was to produce a chitosan/alginate/β-TCP

scaffold by FRM, freeze-drying, and RP techniques, and posteriorly perform a comparative

study.

Page 31: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

14

1.5 Aims

In the present study, the main goal was to study the influence of three different production

techniques on the properties of a chitosan/alginate/β-TCP scaffold, in order to find the most

suitable method to produce this 3D construct for bone regeneration. For that, the objectives

of this work are:

Production of scaffold by Rapid Prototyping;

Production of scaffolds by Freeze-drying;

Production of scaffolds by Foam Replication Method;

Evaluation of the mechanical, physicochemical, and biological properties of the

scaffolds.

Page 32: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

Chapter II

Materials and Methods

Page 33: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

16

2.1 Materials

Amphotericin B, bovine serum albumin (BSA), dulbecco’s modified eagle’s medium (DMEM-

F12), ethanol (EtOH), ethylenediaminetetraacetic acid (EDTA), glutaraldehyde, high

molecular weight chitosan (Mw= 310.000-375.000 g/mol), LB Broth, L-glutamine,

paraformaldehyde (PFA), penicillin G, phosphate-buffered saline (PBS), sodium alginate (Mw=

120.000–190.000 g/mol), sodium tripolyphosphate (TPP), streptomycin, trypan blue, trypsin ,

and 7-hydroxy-3H-phenoxazin-3-one-10-oxide (resazurin) were bought from Sigma-Aldrich

(Sintra, Portugal). β-TCP powder was aquired from Panreac® (Barcelona,Spain). Fetal bovine

serum (FBS) was purchased from Biochrom AG (Berlin, Germany). Human osteoblast cells

(CRL-11372) and Staphylococcus aureus (25923) were obtained from American Type Culture

Collection (VA, USA). Propidium Iodide (PI) was purchased from Invitrogen (Carlsbad, USA).

Acetic acid and sodium hydroxide (NaOH) were bought from Pronalab.

2.2 Methods

2.2.1 Production of the scaffolds

Initially, a solution of Alginate/Chitosan/β-TCP was optimized and finally prepared in a ratio

1/1/8 (% wt/wt/wt). Briefly, chitosan powder was dissolved at 2% (wt/v) in a 1% acetic acid

solution and left overnight, under magnetic stirring. Then, alginate powder was added to the

solution and left stirring for 5 hours. Finally, β-TCP powder was added and the final mixture

was homogenized using a X10/25 ultraturrax (Ystral, Germany). After dissolution of all the

components, the solution was processed with the different techniques.

2.2.1.1 Production of the scaffolds by the Foam Replication Method (FRM)

To prepare the scaffolds by FRM, polyurethane (PU) foam was cut into cubes (1 x 1 x 1 cm) to

be used as sacrificial template [63]. These were then washed with ultrapure water, and left

to dry at RT. Then, the previously prepared solution was impregnated in the PU foams. The

wet samples were subsequently immersed in a 6% TPP (wt/v) solution, for 24 h, and dried at

room temperature (RT). Finally, the samples were step-wise heated in a furnace at 1ºC/min

to 1100ºC, and then kept at this temperature for 240 min [66].

Page 34: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

17

2.2.1.2 Production of the scaffolds by the Freeze-Drying technique

The scaffolds were also prepared by freeze-drying [57, 90, 91]. Briefly, the previously

produced solution was deposited into a 48-well polystyrene plate, frozen at -20ºC for 24 h,

and then lyophilized for another 24 h. Thereafter, the lyophilized 3D structures were

removed from the mold and cross-linked with a 6% TPP (wt/v) solution for 24 h. Finally, the

scaffolds underwent another similar cycle of freezing and lyophilization.

2.2.1.3 Production of the scaffolds by Rapid Prototyping

To prepare the scaffolds by RP, the 3D structures were first designed in Solidworks® software

(Dassault Systèmes, Massachusetts, USA). This software allows the creation of the 3D scaffold

models with controllable structure. A Fab@home 3D printer was chosen to produce the

scaffolds, as previously described by Diogo et al [58]. In order to print the scaffolds, the file

was converted and stored in stereolithography (STL) format. A 10cc disposable syringe barrel

was filled with the previously prepared solution, and the solution was extruded trough a 20G

polypropylene tapered nozzle. Thereafter, the produced scaffolds were immersed in a 6 %

TPP (wt/v) solution for 24 h. Subsequently, the scaffolds were frozen at -80ºC and lyophilized

for 24 h, in order to allow their drying.

2.2.2 Morphological and physicochemical characterization of the scaffolds

2.2.2.1 Morphological analysis of the scaffolds

Scanning electronic microscopy (SEM) analysis was performed in order to assess the

morphology and pore sizes of the different scaffolds. This analysis was performed by adapting

the method reported by Valente et al [28]. The scaffolds were mounted on aluminum stubs

using double-sided adhesive tape, and sputter coated with gold using a Q150R ES (United

Kingdom) sputter coater. The materials were analyzed using a Hitachi S-3400N (Tokyo, Japan)

scanning electron microscope operated at an acceleration voltage of 20 kV, at various

magnifications.

2.2.2.2 Fourier Transform Infrared Spectroscopy analysis

Fourier-transform infrared (FTIR) spectroscopy was used to analyze the physicochemical

characteristics of the scaffolds [38, 64]. Briefly, all samples were crushed, the resulting

powders were mounted on a diamond window, and the spectra recorded on a Nicolet iS10

FTIR spectrophotometer (Thermo Scientific, USA). The individual powder components used in

the production of the scaffolds were also analyzed [55]. FTIR spectra represent the average

Page 35: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

18

of 250 scans, between 500 and 4000 cm-1, at a resolution of 4 cm-1. The acquired data was

then processed using Omnic Spectra analysis software.

2.2.2.3 X-Ray diffraction analysis

To evaluate the crystallinity and characteristic phases of the scaffolds, X-ray diffraction (XRD)

measurements were performed with a X-ray diffractometer (Rigaku Americas Corporation,

USA). XRD technique was used to assess the material composition and internal structure, and

to confirm the presence of β-TCP in the scaffolds. Samples were mounted in silica supports

and the data was recorded over a range of 5º to 90º 2θ degrees, with continuous scans at a

rate of 1º/min with a copper ray tube operated at 30 kV and 20 m [58].

2.2.2.4 Energy dispersive spectroscopy

In order to assess the elemental composition of the scaffolds, Energy dispersive spectroscopy

analysis (EDS) (Rontec) was also performed. Prior to data acquisition, the scaffolds were cut

into slices, mounted in aluminum stub supports and air-dried at RT [58, 63].

2.2.3 Water contact angle measurement

In order to analyze the scaffolds hydrophilicity, their water contact angle was measured. The

assay was performed using a sessile drop method, adapted from Diogo et al, with water as

reference fluid [58]. The contact angle data were acquired in a Data Physics Contact Angle

System OCAH 200 apparatus, operating in static mode at RT. For each sample, water drops

were placed at different locations of the analyzed scaffolds.

2.2.4 Total porosity evaluation

In order to assess the porosity (P) of the scaffolds, the method previously described by Nie et

al was adapted [92]. The porosity of each scaffold was determined through the total amount

of absolute ethanol (EtOH) that the scaffolds were able to absorb in 48 h, using equation 1

[93].

Page 36: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

19

Where W1 the weight of the dry scaffold and W2 is weight of the wet scaffold, dethanol is the

density of the ethanol at RT and Vscaffold is the volume of the wet scaffold. For each scaffold,

three replicates were analyzed and data represent the average of each replicate.

2.2.5 Swelling capacity of the scaffolds

The swelling capacity of scaffolds (S) was determined by following a method adapted from

Coimbra et al [94]. In brief, the different scaffolds were placed in falcons containing Tris

buffer (pH=7.4), at 37ºC. At predetermined intervals, the swollen samples were removed

from the solution, the excess of Tris was removed with filter paper and the samples were

weighted. After, the scaffolds were re-immersed into the swelling medium. The scaffolds

were measured in triplicate, and the swelling ratio was determined by using equation 2:

Where Ws is the final weight of scaffolds and Wd is the initial weight of scaffolds.

2.2.6 Degradation of scaffolds

Scaffolds degradation was also assessed by adapting a method described by Christopher et al.

[95]. Briefly, the scaffolds were weighted and placed in 6 well-plates, containing incomplete

Dulbecco’s modified Eagle medium (DMEM-F12) at 37ºC, for 4 weeks. Once a week, they were

removed from the medium, washed with ultrapure water, dried and weighted. Four replicates

of each type of scaffold were analyzed. The scaffolds mass loss was determined through

equation 3 [95]:

Where Mf is the final mass of the scaffolds and Mi is the initial mass of the scaffolds.

Page 37: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

20

2.2.7 Mechanical characterization of the different scaffolds

In order to assess the mechanical behavior of the different scaffolds, their resistance to

compression and Young´s Modulus were analyzed. The assays were performed by following a

method adapted from Santos et al [64]. Briefly, the produced scaffolds were first cut into

similar fragments and measured. Compression assays were then performed using a Zwick®

1435 Material prüfung (Ulm, Germany), with a crosshead speed of 0.2mm/min and a load cell

of 5 kN. Six samples of each scaffold were tested.

The determination of the resistance to compression (Ts) of each scaffold was determined

through equation 4 [72]:

Where F is the load at the time of the fracture and a and l represent the width and length of

the scaffold, respectively.

Young’s Modulus (YM) was obtained from the stress-strain relations calculated by applying the

equation 5 [96]:

)

Where Hd is the scaffold height deformation and Ts is the scaffold tensile strength. Average

values and standard deviations (s.d.) were determined for each sample.

2.2.8 Biological characterization of the chitosan/alginate/β-TCP scaffolds

2.2.8.1 Characterization of the cytotoxic profile of the scaffolds

In order to study the biocompatibility of the scaffolds, human osteoblast cells (CRL-11372)

(hOB) were used. First, hOB were cultured in DMEM-F12, containing 10% of heat inactivated

fetal bovine serum (FBS), streptomycin (100 μg /mL) and gentamicin (100 μg /mL) antibiotics,

in 75 cm2 T-flasks. Cultures were maintained in a humidified environment at 37ºC, with 5%

CO2, until confluence was attained. Then, cells were trypsinized with 0.18 % trypsin (1:250)

and 5 mM EDTA, centrifuged at 260rpm for 5 min, and the resulting pellet was resuspended in

5 mL of medium. After that, cellular density (number of cells per mL) was determined by

using the trypan blue method, according to equation 6:

Page 38: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

21

Before being placed in contact with the cells, the scaffolds were cut into the desired size,

and placed into 48-well plates. They were subsequently sterilized under ultraviolet (UV) light

for 1 h [49]. The scaffolds were neutralized with NaOH 0.1M for 1 h, and thereafter washed

with incomplete medium at 37 ºC, several times [97]. Finally, cells were seeded in contact

with scaffolds at density of 10x103 cells per well, to evaluated cell viability and proliferation

at days 1, 4 and 7. The culture medium was changed every two days until day 7 [49, 58].

In order to evaluate the cytotoxicity of the scaffolds, after the determined incubation

periods, resazurin assays were performed [98]. Briefly, the medium was removed from each

well and replaced by a mixture of 10% (v/v) of resazurin and 90% (v/v) of medium, and the

cells were incubated for 4 h, at 37ºC, under a 5% CO2 humidified atmosphere. The

fluorescence was measured using a plate reader spectrofluorometer (Spectramax Gemini XS,

Molecular Devices LLC, USA) at an excitation/emission wavelength of 560/590 nm,

respectively. EtOH treated cells were used as positive controls (K+) while cells without

samples were used as negative controls (K-).

2.2.8.2 Proliferation of the osteoblast cells in the presence of the scaffolds

After the sterilization and washing of the scaffolds, hOB were also seeded at a density of

10x103 cells per well, on the surface of the materials for 1, 4 and 7 days, as previously

described. Cell growth was monitored using an Olympus CX41 inverted light microscope

(Tokyo, Japan) equipped with an Olympus SP-500 UZ digital camera [58].

Cellular attachment and morphology were also characterized by SEM analysis. After 1 and 7

days of seeding with hOB, the culture medium was removed and the samples were washed

with PBS and fixed overnight with 2.5% glutaraldehyde (v/v) at 4°C. Samples were rinsed with

PBS and dehydrated with graded EtOH solutions (70, 80, 90 and 100%). Posteriorly, the

samples were frozen with liquid nitrogen and freeze-dried for 3 h. Then, they were mounted

on aluminium stubs using double sided adhesive tape, and sputter-coated with gold using a

Quorum Q150R ES sputter coater. Finally, SEM images were acquired with an acceleration

voltage of 20 kV, at variable magnifications, using a Hitachi S-3400N scanning electron

microscope [99].

Page 39: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

22

2.2.8.3 Confocal Laser Scanning Microscopy

In order to analyse the cell distribution within the 3D scaffolds, Confocal Laser Scanning

Microscopy (CLSM) analysis was performed by adapting a protocol described by Miguel et al

[49]. In brief, 4 days after cell seeding, the culture medium was removed and the cells were

washed with PBS at RT. The remaining cells were then fixed with 4% PFA for 15 min at RT,

and then rinsed three times with PBS. Afterwards, the cell nucleus was labelled with

propidium iodide (PI) (15 mM) during 15 min, followed by five additional washes with PBS.

The cell-seeded scaffolds were then transferred into μ-Slide 8 well Ibidi® chamber coverslips

and imaged in a Zeiss LSM710 confocal microscope (Carl Zeiss SMT Inc., USA) equipped with a

Plan-Neofluar 10×/NA 0.3 and Plan-Apochromat 40×/1.4 Oil DIC objectives. All data was

acquired in z-stack mode with a step of 4.67 μm. Z-stacks were then rendered into 3D images

using Zeiss LSM 710 software. Depth code rendering of z-stacks were also performed using

Zeiss software with the open GL and transparent rendering mode to provide visualization of

cell spatial distribution within the scaffold architecture.

2.2.9 Evaluation of the antibacterial activity of the scaffolds

To evaluate the antimicrobial activity of the scaffolds, Staphyloccocus aureus was used as a

model of Gram-positive bacteria. The bacterial growth in the presence of the scaffolds was

assessed through a Kirby-Bauer disk diffusion method, as previously described by Kim et al

[100]. The scaffolds were placed on the surface of a plate of LB agar, in contact with

Staphylococcus aureus at a concentration of 0.5x10⁸ CFU/ml. Then the petri plates were

incubated for 24h at 37°C [100]. Finally, the antibacterial activity of the scaffolds was

assessed through the observation of the inhibition hales around the samples.

2.2.10 Statistical Analysis

Statistical analysis of the obtained results was performed using one-way ANOVA with

Dunnet´s post hoc test and Newman-Keuls test. A p value lower than 0.05 (p<0.05) was

defined as statistically significant. The obtained results were expressed as the mean ±

standard error of the mean of at least three independent experiments.

Page 40: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

Chapter III

Results and Discussion

Page 41: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

24

3.1 Macroscopic properties of the scaffolds

In this study, chitosan/alginate/β-TCP scaffolds were produced by three different production

techniques (FRM, Freeze-drying and RP) and their properties were characterized. The

comparison of the scaffolds properties allows the comprehension of how the different

manufacturing processes can influence the suitability of these scaffolds for bone tissue

regeneration [36, 70]. The chosen production techniques were selected since they are

commonly used for the production of bone scaffolds. The polymeric materials used for the

production of the scaffolds were chosen due to their ability to mimic the bone matrix.

Chitosan and alginate are natural polymers that present biocompatibility, biodegradability

and low toxicity properties. Moreover, chitosan has bactericidal activity [43, 51]. The

bioactive ceramic β-TCP presents a similar composition to the inorganic components of the

bone matrix, possessing osteoindutivity and osteocondutivity, biocompatibility and adequate

biodegradability. Furthermore, this ceramic improves the mechanical resistance of the

scaffolds, which is extremely important for bone TE applications [65, 101].

Figure 8 presents the macroscopic images of the produced scaffolds. The scaffolds produced

by RP (figure 8 A) presented a more complex and controlled macroporous structure, that is

important for the promotion of cell adhesion and capillary in-growth [69, 102]. The RP

method allows the control of all the structural parameters, thus leading to a scaffold with the

desired shape [103]. In opposition, the scaffolds produced by Freeze-drying (figure 8 B) and

FRM (figure 8 C) presented structures limited by the templates used. For instance, the

samples produced by Freeze-drying were limited to their cylindrical molds [57, 86] while the

FRM samples were limited to the cubic template, which was sacrificed during sintering [63].

Another important property of the surface of scaffolds is roughness, since a rough surface

increases the contact areas between the materials and the cells, allowing better cell

attachment, proliferation and differentiation [13, 69]. Moreover, it is described that OB are

able to attach more rapidly and synthesize more extracellular matrix in the presence of

rougher surfaces [22, 31].

The scaffolds produced by RP and Freeze-drying showed smoother and soft surfaces, while

scaffolds produced by FRM presented more rough surfaces (Figure 8 C, C1).

Page 42: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

25

Figure 8: Macroscopic images of the final scaffolds produced by RP (A), freeze-drying (B), and RP(C).

SEM images were also acquired in order to characterize the internal structure and pore sizes

of the scaffolds (figure 9). It has been previously reported that porous structures with pore

sizes above 100 µm allow cell penetration and proliferation inside the scaffolds, as well as the

exchange of gases and nutrients [28, 29, 69].

Through the analysis of the SEM images (figure 9), it is clear that each scaffold presents a

different internal structure and pore size. The RP samples (figure 9 A and B) showed a highly

porous and interconnected structure, with pore sizes around 150 µm, a suitable size for bone

regeneration [69]. Such porosity might be attributed to the liophilization process used in the

production of this scaffold [37]. The images of the Freeze-drying samples (figure 9 C and D)

showed that these scaffolds had a porous structure, presenting pore sizes around 220 µm.

These characteristics provide a suitable environment for cell growth and proliferation [69]. It

is important to notice that these scaffolds presented larger pores than the RP samples. As

proposed by Kang et al, the different freezing temperatures (-20ºC and -80ºC) that these

scaffolds were exposed, lead to different pore sizes within the scaffolds structure. In brief,

the Freeze-drying samples were frozen at -20ºC leading to a slower freezing and higher

crystals formation. On the other hand, RP scaffolds were frozen at -80ºC leading to faster

freezing and smaller crystals. Consequently, after the liophilization step, the respective

crystals in the scaffolds structure were removed, creating different pore sizes [91].

The FRM samples (figure 9 E and F) also revealed a highly porous structure with large pores,

similar to those of the Freeze-drying samples. These results can be attributed to the sintering

process that led to the loss of the mold and polymers in the sample [104, 105], remaining just

the ceramic. Such loss can negatively influence the scaffold role in bone regeneration, since

Page 43: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

26

the polymers would provide additional desirable properties to the biomaterial, such as

antibacterial activity and swelling capacity. Tampieri et al described the use of this

technique for the production of similar porous structures, with pore sizes between 200 and

500 µm [76], which are comparable to those obtained here.

Figure 9: SEM images of the internal porosity of the scaffolds produced by RP (A, B), Freeze-drying (C,

D), and FRM (E, F) techniques.

Page 44: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

27

3.2 Physicochemical characterization of the scaffolds

3.2.1 Fourier Transform Infrared Spectroscopy

FTIR analysis of the biomaterials was performed, including comparison of each scaffold

spectrum with the raw powders spectra (figure 10), since it is important to evaluate if the

different techniques used for scaffolds production have any influence on their composition.

Figure 10: FTIR analysis of the 3D scaffolds produced by the three different technics and comparing

with the powders.

The FTIR spectra of alginate, chitosan and β-TCP were obtained from the respective powders.

Through the analysis of these spectra (figure 10), it is possible to observe that alginate

presented bands at 1421 cm-1, 1620 cm-1 and 3429 cm-1, corresponding to the COOH stretch

and C=O stretch of the alginate carboxylic groups, and to a O-H stretch vibration, respectively

[41]. Chitosan spectrum showed peaks at 1157 cm-1, 1422 cm-1 and 1620 cm-1 that were

assigned to the N-H stretch of the amine, C=O stretch and N-H stretch of the amide,

respectively [55]. Moreover, the large band between 3000 cm-1 and 3500 cm-1, present in the

Page 45: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

28

polymers, belonged to the O-H stretch of water [41, 55]. β-TCP spectrum revealed a major

band at 1020 cm-1 that was assigned to the P-O stretching [58, 67].

The spectra of the scaffolds produced by RP and Freeze-Drying, showed the characteristic

peaks of chitosan, alginate and β-TCP powders, suggesting that these components were

present in the scaffolds and that the composition of the scaffolds was not altered when

different techniques were used for scaffolds production. However, the spectrum of the

scaffolds produced by FRM revealed only the characteristic peak of β-TCP, suggesting that the

scaffolds lost its polymeric content. This can be explained by the fact that the exposition of

the scaffolds to a sintering temperature of 1100ºC led to a removal of chitosan and alginate

contents [105-107]. In fact, for temperatures higher than 800ºC, the polysaccharides are

desintegrated, resulting in a scaffold composed only by the ceramic [104-107].

3.2.2 X-Ray diffraction analysis (XRD)

In order to confirm the presence of crystalline phases of β-TCP in the scaffolds, XRD analysis

was also performed. Figure 11 shows that the commercial β-TCP powder presented two

intense peaks located at 2θ=25º-35º. Such peaks were also observed in all the manufactured

scaffolds. The absence of anomalous crystalline phases suggested that the ceramic

component of the scaffolds preserved its crystalline structure [58]. Furthermore, the

intensified peaks present in the scaffold produced by FRM can be explained by the sintering

process that only left β-TCP in the samples [60, 100]. In addition, Sanosh et al reported that

when the sintering temperature increase, the β-TCP peaks of the sample become sharper,

representing a better crystallinity [78]. Furthermore, it is reported that β-TCP is stable below

1180ºC and when the sintering temperature reaches 1200ºC β-TCP acquires its α-TCP

conformation [60]. Nevertheless, the sintering of β-TCP at 1100ºC does not affect its beta

conformation, as previously demonstrated by Chen et al [66].

Page 46: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

29

Figure 11: X-Ray spectra of β-TCP and 3D scaffolds. Peaks inside the rectangle represent the

characteristic peaks of the ceramic present in the scaffolds.

3.2.3 Energy Dispersive Spectroscopy (EDS)

An EDS analysis was also performed to evaluate the elemental composition of the scaffolds

(table 2). The carbon (C) content of the scaffolds is provided by the polymers in the samples.

The results showed that the amount of C was higher in the scaffolds produced by RP and

Freeze-drying, when compared to the ones produced by FRM. This result is in accordance with

the FTIR results, since the polymers disintegrated during sintering [104, 105]. In addition, the

calcium (Ca) and phosphate (P) content were similar between the samples, demonstrating

that the ceramic component (figure 8) was not affected, as observed in the FTIR and XRD

results. Sodium (Na) and oxygen (O) percentages were also similar between the scaffolds,

showing that no additional changes occurred in the scaffolds.

Page 47: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

30

Table 2: EDS analysis of the produced 3D scaffolds. The atomic percentage (At.%) of the elements in the

different scaffolds is evaluated.

At.%

RP

Freeze-drying

FRM

C

16.98%

16.64%

7.03%

Ca

10.31%

12.97%

14.28%

P

10.04%

10.19%

12.20%

Na

7.21%

5.33%

7.91%

O

55.46%

54.87%

58.59%

FTIR, XRD and EDS results are important to analyze how the different production techniques

influenced the physicochemical properties of the chitosan/alginate/β-TCP scaffolds. The

major modifications in the scaffolds composition were found in the ones that were produced

by FRM. In fact, several studies reported the use of FRM only for the production of ceramic

scaffolds [60, 63, 64]. Therefore, this technique should only be applied when the goal is to

produce ceramic 3D constructs. Nevertheless, it is reported that β-TCP scaffolds can also

have suitable properties to be used in bone TE, including good biocompatibility and

biodegradability [66, 67, 108].

3.3 Porosity of scaffolds

Scaffold porosity is important for cell migration and nutrients exchange [27]. Moreover, the

porosity directly influences the mechanical strength of the scaffolds. Excessive porosity will

cause a decrease in the mechanical strength, and the consequent disintegration of the

materials. On the other hand, scaffolds presenting lower porosity will have a higher strength

[69]. In order to analyze the porosity of the scaffolds, a liquid displacement method, using

ethanol, was performed [8].

The results showed that all the scaffolds presented low porosity values (figure 12), with the

Freeze-Drying samples presenting the highest porosity values (12.6%), and the RP samples

showing the lowest (5.8%). As explained above, this difference in the results comes from the

freezing process. Freeze-drying samples were frozen at -20ºC, leading to a slower freezing

and higher pores after liophilization, while RP scaffolds were frozen at -80ºC, leading to a

Page 48: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

31

faster freezing and smaller pores [91]. FRM samples presented almost as much porosity as the

Freeze-drying samples (10.6%). This can be explained by the fact that these samples were

sintered and only contain β-TCP, leading to an internal structure with large empty spaces.

These results are in accordance with the data obtained through SEM analysis (figure 9). These

porosity results showed that all scaffolds have suitable porosities for mimicking the cortical

bone porosity (2-13%) and acceptable pore sizes for cellular internalization [69].

Figure 12: Total porosity the different scaffolds.

3.4 Contact angle determination

Along with porosity, the hydrophilicity of a scaffold can interfere with its bioactivity and

regenerative capacity. Hydrophilic scaffolds enable the adherence and proliferation of cells

at its surface [38]. Thus, the contact angle of the scaffolds was determined in order to

analyze the hydrophilicity of the scaffolds. A scaffold is considered hydrophilic when the

water contact angle is below 90º. [58].

Page 49: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

32

The obtained results (table 3) showed that the RP and Freeze-drying samples presented good

surface hydrophilicity. The hydrophilic character in these samples is mostly a consequence of

the high carboxyl and hydroxyl groups present in the polymers [28]. The FRM sample

presented similar contact angle values to the other samples. Although FRM samples do not

have any polymeric content in its structure, as demonstrated by the FTIR and EDS analysis, it

is reported in several studies that β-TCP has hydrophilic character, thus confering

hydrophilicity to this scaffold [58, 60, 64, 109]. These results suggest that all the produced

scaffolds present suitable hydrophilicity for bone regeneration applications. Studies

performed by Valente et al reported that hydrophilic scaffolds allow cell adhesion and

differentiation as well as oxygen and nutrients exchange from body fluids into the scaffolds,

which have an essential role for the successful repair of bone defects [28].

Table 3: Water contact angle values determined for the different scaffolds.

Scaffolds

Water contact angle

RP

25.30 ± 6º

Freeze-drying

32.13 ± 6º

FRM

27.95 ± 6º

3.5 Swelling capacity of the scaffolds

Swelling capacity is an important factor to consider in a scaffold for bone tissue engineering

[23, 29, 32]. It refers to the scaffold’s ability of fluid uptake under physiological conditions.

This fluid diffusion is crucial for the internalization of essential nutrients, cells and bioactive

molecules within the scaffold [59]. Nonetheless, excessive swelling can lead to implant

loosening and decreased mechanical properties [59]. In order to evaluate the scaffold’s

ability to absorb body fluids, swelling studies were performed [49].

The obtained results showed that all scaffolds presented different swelling behaviors. Freeze-

drying and RP samples revealed fast swelling in the first 30 minutes. According to Morais et

al, the osmotic pressure is greater than the forces of the crosslinking bonds that stabilize the

structure of the polymers network, leading to a rapid water uptake. After the void regions of

this network are filled, an equilibrium state is achieved [43]. Herein, this equilibrium is

observed after 1 h (figure 13). In addition, Patel et al suggested that the swelling behavior of

scaffolds containing chitosan is proportionally dependent to the biomaterials porosity. Such

behavior is observed between the RP and Freeze-drying samples. The scaffolds produced by

Freeze-drying have higher porosity and swelling behavior than those produced by RP. FRM

Page 50: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

33

samples presented the lowest swelling values (about 25%), stabilizing immediately after

(figure 13). As described before, this scaffold it is mainly composed by a highly porous

ceramic structure. According to Morais et al, the higher the polymer concentration within a

scaffold, the higher is its swelling behavior [43]. Therefore, the low swelling degree

presented by the FRM samples is resultant from its composition.

Figure 13: Swelling behavior of the different scaffolds.

3.6 Characterization of the degradation profile of the scaffolds

A scaffold’s biodegradation is essential to promote a controlled bone regeneration [32]. An

uncontrolled degradation would result in biomaterial disintegration, hindering the bone

regeneration process. Contrarily, if there is insufficient degradation, it can lead to

deformities during bone remodeling. As such, the degradation rate should match the

formation rate of new tissue to serve as useful template [62, 108]. The degradation of the

scaffolds was analyzed for 4 weeks, in contact with culture medium (figure 14) [95].

According to Holzapfel et al, the degradation rate of a scaffold decreases with decreasing

porosity [32]. This is in accordance with the results obtained for the swelling behavior of the

Freeze-drying and RP samples. The balance between water diffusion into the polymeric

material and the kinetics of hydrolytic polymer chain degradation determine the degradation

Page 51: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

34

rate of the scaffolds [32]. In the literature it is reported that alginate and chitosan are

degraded through hydrolysis of glycosidic bonds, which leads to a weight decrease with time

[43]. Nevertheless, the crosslinking of alginate with chitosan can lead to greater scaffold

stability, leading to a slower degradation rate [59].

Figure 14: Degradation rate of the different scaffolds.

FRM samples presented the fastest degradation rate, with a matter loss of more than 60%

after 4 weeks, that is in accordance with what was previously reported by Sowjanya et al

[59]. β-TCP is known for having higher biodegradability than other ceramics, like HA [63, 64].

Furthermore, the porosity of these scaffolds can also contribute to a higher degradation rate

[32].

In a study performed by Carano et al, the author reports that it takes more than three weeks

for a human bone fracture to start healing [110]. Consequently, it is essential that a scaffold

aimed to improve this process possesses an adequate degradation time. With this degradation

results it is possible to observe that the scaffolds that would have the best degradation

behavior to be applied in bone regeneration are those produced by RP and Freeze-drying

techniques.

Page 52: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

35

3.7 Mechanical characterization of the scaffolds

The mechanical behavior of the 3D scaffolds was characterized by Compressive strength and

Young´s modulus assays (figures 15 and 16). A suitable scaffold to be used for bone

regeneration must have both flexibility and resistance to compression. These characteristics

have profound effects on load bearing and cellular behavior during bone regeneration [29].

Through the analysis of figures 15 and 16, it is possible to observe that the RP samples

presented low compressive strength values (0.3 MPa), and higher Young modulus values (30

MPa). The same behavior was observed in the Freeze-drying samples that presented

compressive strength values (0.6 MPa) lower than Young modulus values (42 MPa). FRM

samples presented the higher compressive strength (1.8 MPa) and the lower Young modulus

(13.8 MPa). Similar behaviors have been previously reported in scaffolds that only contain β-

TCP [111]. Furthermore, FRM samples presented a high variability in compressive strength.

This is due to the fact that these scaffolds are composed mainly by β-TCP, that despite being

known for its resistance [62], is also reported as a brittle material, with a low fatigue

resistance [32, 60, 64].

These results showed that FRM scaffolds have values of compressive strength closer to that of

human trabecular bone, but with limited flexibility, as noted by its Young modulus. This

suggests that they can only be used for non load bearing applications [7].

Page 53: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

36

Figure 15: Characterization of the compressive strength of the different scaffolds. Statistical analysis

was performed using one-way ANOVA with Newman-Keuls test (* p < 0.05).

Figure 16: Characterization of scaffolds Young’s Modulus. Statistical analysis was performed using one-

way ANOVA with Newman-Keuls test (* p < 0.05).

Page 54: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

37

3.8 Biological characterization of the scaffolds

3.8.1 Characterization of the cytotoxic profile of the scaffolds

In order to evaluate the cytotoxicity of the chitosan/Alginate/β-TCP scaffolds after 1, 4 and 7

days, in vitro resazurin assays were performed (figure 17) [98]. It is possible to observe that

there is a significant difference between the positive control (K+) and the negative control (K-

). This suggests that the biomaterials did not affect cell viability during the incubation period.

Figure 17: Evaluation of the osteoblasts viability through a resazurin assay, after 1, 4 and 7 days of

incubation in the presence of the scaffolds. Positive control (K+) and negative control (K-), indicate

dead and viable cells in the absence of scaffolds, respectively. Each result is the mean ± standard error

of the mean of at least three independent experiments. Statistical analysis was performed using one-

way ANOVA with Dunnet’s post hoc test (* p < 0.05).

It is possible to observe that there was an increase in cell viability from 1 to 7 days when they

were seeded in contact with scaffolds, suggesting that these 3D constructs have suitable

properties for cell growth and proliferation. No significant differences were noticed between

the different scaffolds. Optical microscopic images of the cells in contact with the scaffolds

are presented in figure 18. These images show that hOB adhered and proliferated, in the

Page 55: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

38

presence of all the scaffolds, in a similar way to that of the negative control (K-). Regarding

the positive control (K+), no cell adhesion or proliferation was observed, and dead cells in

their typical spherical shape were observed. Such results show that the different

manufacturing techniques did not affect the biocompatibility of the scaffolds.

Page 56: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

39

Figure 18: Macroscopic images of human osteoblast cells in the presence of the scaffolds produced by

RP (A, B, C), Freeze-drying (D, E, F) and FRM (G, H, I), after 1, 4 and 7 days of seeding. Negative control

(K-) is represented in images J, K and L, and positive control (K+) is represented in images M, N and O.

Page 57: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

40

3.8.2 SEM and CLSM images

In order to characterize cellular adhesion and proliferation, both on the surface and inside

the scaffolds, SEM and CLSM analysis were acquired (figure 19 and 20). SEM images showed

that cells adhered and spread within the different scaffolds after 1 day (figure 19 A, C and E).

It is possible to observe that after 7 days of incubation (figure 19 B, D and F), the cells

proliferated throughout the scaffold’s surface, showing that all scaffolds presented suitable

biological properties.

Figure 19: SEM images of hOB morphology after being cultured for 1 and 7 days in contact with the

different scaffolds.

To further characterize cell internalization within the different scaffolds, CLSM images were

also acquired (figure 20). hOB were seeded in contact with the different scaffolds for four

days, and labeled with PI. Through the analysis of the images 20 A, C and E, it can be

observed that cells grew inside the porous structure of the scaffolds. The images also give an

Page 58: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

41

idea of the cells location in the scaffolds, demonstrating that the biomaterials have suitable

porosity to accommodate hOB. Figures 20 B, D and F represent the color depth analysis of the

different scaffolds. The images show that hOB are capable of migrating and attaching into

deep sections of the scaffolds, with most cells localized up to 20-40 µm within the scaffolds

structure. This is very important for a further deposition of bone matrix inside the scaffold

and posterior remodeling of the bone defect [69]. These results suggest that all the scaffolds

presented suitable structures for allowing cell internalization, which is crucial for bone

regeneration.

Figure 20: CLSM images of the scaffolds produced by RP, Freeze-drying, and FRM, seeded with OB after

4 days of incubation. Images A, C and E show the cell distribution on the surface of scaffolds and the

respective representative orthogonal images of the x-z and y-z planes. Images B, D and F show depth

colour coding of the respective scaffolds, obtained in the same period. White bars represent 200 μm.

Page 59: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

42

3.9 Antibacterial activity of the scaffolds

It is well known that in the surgery field orthopaedic interventions have a high risk of

triggering inflammatory responses due to bacterial colonization of the biomaterial’s surface

[112]. Therefore, the antibacterial activity of the scaffolds against S.aureus was also

evaluated through a Kirby-Bauer disk diffusion method [100].

Chitosan, due to its protonated amine groups, can interact with the electronegative residues

present at bacteria surface, decreasing their cell wall permeability and preventing nutrients

from entering the cell, thus resulting in bacteria death [46, 49]. The results obtained in the

studies of antibacterial activity revealed that, apart from the scaffolds produced by FRM, all

the scaffolds inhibited S.aureus growth, inducing an inhibitory halo around the scaffolds

(Figure 21). As previously demonstrated, the FRM samples do not possess polymeric content,

and thus did not present any antibacterial activity [105]. This antibacterial test of the

scaffolds against S.Aureus suggest that the intrinsic characteristics of the RP and Freeze-

drying samples can contribute to limit the risk of infections [49].

Figure 21: Scaffolds antibacterial activity trough a Kirby-Bauer diffusion method, where inhibition halos

around RP (A), Freeze-drying (B) and FRM (C) samples were evaluated using S.Aureus.

3.10 Summary of the properties of the produced scaffolds

Table 4 presents a summary of the results obtained, emphasizing the different properties for

each scaffold. Analyzing the table, it is possible to conclude that the scaffold that has the

best properties to be used in bone TE, was produced by Rapid prototyping. In fact, several

studies state that this technique emerged in bone TE to allow the control of the morphology

Page 60: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

43

and porosity [26, 58, 85, 102]. Herein, scaffolds produced with RP also showed the best

properties.

Table 4: Summary of the properties of the chitosan/alginate/β-TCP scaffolds produced in this study by

RP, Freeze-drying and FRM techniques. Qualitative grading of scaffolds properties goes from less

suitable (+) to more suitable (+++) properties for the scaffolds to be used in bone tissue engineering.

Page 61: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

Chapter IV

Conclusions and Future Perspectives

Page 62: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

45

4.1. Conclusions and future perspectives

Bone defects are an increasing problem, affecting millions of people worldwide. Temporary

and biodegradable scaffolds are a good choice to overcome this health problem. As such,

choosing the ideal manufacturing technique to produce these scaffolds is essential to control

the final biological, physicochemical and mechanical properties of the 3D constructs.

Nowadays, several techniques are used for the production of scaffolds. In this study,

chitosan/Alginate/β-TCP scaffolds were produced by three different techniques (Rapid

prototyping, Freeze-drying and Foam replication method), in order to analyze the influence

of the manufacturing processes on scaffold’s properties. The scaffolds produced by Rapid

prototyping presented a controlled shape, as well as a desirable macroporosity. Porosity,

hydrophilicity and swelling analysis demonstrated that the scaffolds produced by Rapid

prototyping and Freeze-drying possessed the most suitable properties for cell internalization

and proliferation. Biocompatibility results revealed that none of the scaffolds presented

cytotoxicity towards hOB, and that these cells were able to adhere and proliferated on the

material’s surface. Nevertheless, the scaffolds produced by Foam replication method lost

their polymeric content during the production process, hindering their biological properties.

Regarding to mechanical resistance, the scaffold that presented the best resistance to

compression was produced by Foam replication method, while the worst resistance values

belonged to the scaffold produced by Rapid prototyping. Overall, the chitosan/alginate/TCP

scaffolds produced by Rapid prototyping combined the best properties to be used in bone TE.

These results confirm that Rapid prototyping can overcome some drawbacks of the other

techniques such as the lack of control in the size and shape of the scaffolds. A macroporous

structure is essential for a better cell internalization and proliferation within a scaffold,

allowing an enhanced bone regeneration to occur. However, the mechanical resistance of this

scaffold is far from the “ideal” values. In a near future, the mechanical properties of the

chitosan/alginate/β-TCP must be improved. The addition of another ceramic compound, such

as hydroxyapatite, would improve the mechanical properties of the scaffolds.

Page 63: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

Chapter V

Bibliography

Page 64: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

47

1. Tate, P. and R.R. Seeley, Seeley's principles of anatomy & physiology. 2009: McGraw-

Hill.

2. Van De Graaff, K.M., Human anatomy. 2002: McGraw-Hill New York.

3. Marieb, E.N. and K. Hoehn, Human anatomy & physiology. 2007: Pearson Education.

4. Sathyendra, V. and M. Darowish, Basic science of bone healing. Hand clinics, 2013.

29(4): p. 473-481.

5. Ralston, S.H., Bone structure and metabolism. Medicine, 2013. 41(10): p. 581-585.

6. Stevens, M.M., Biomaterials for bone tissue engineering. Materials today, 2008. 11(5):

p. 18-25.

7. Lamya Karim, et al., The mechanical behaviour of bone. Osteoporosis, 2013. 1: p.

431-452.

8. Salgado, A.J., O.P. Coutinho, and R.L. Reis, Bone tissue engineering: state of the art

and future trends. Macromolecular bioscience, 2004. 4(8): p. 743-765.

9. Martin, T.J., K.W. Ng, and N.A. Sims, Basic principles of bone cell biology.

Translational Endocrinology of Bone: Reproduction, Metabolism, and the Central

Nervous System, 2012. 1: p. 5-26.

10. Jayakumar, P. and L. Di Silvio, Osteoblasts in bone tissue engineering. Proceedings of

the Institution of Mechanical Engineers, Part H: Journal of Engineering in Medicine,

2010. 224(12): p. 1415-1440.

11. Bonewald, L.F., The amazing osteocyte. Journal of Bone and Mineral Research, 2011.

26(2): p. 229-238.

12. Jang, J.-H., O. Castano, and H.-W. Kim, Electrospun materials as potential platforms

for bone tissue engineering. Advanced drug delivery reviews, 2009. 61(12): p. 1065-

1083.

13. Martin, T. and E. Seeman, New mechanisms and targets in the treatment of bone

fragility. Clinical science, 2007. 112: p. 77-91.

14. Sommerfeldt, D. and C. Rubin, Biology of bone and how it orchestrates the form and

function of the skeleton. European Spine Journal, 2001. 10(2): p. S86-S95.

Page 65: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

48

15. Vaananen, H., et al., The cell biology of osteoclast function. Journal of cell science,

2000. 113(3): p. 377-381.

16. Raggatt, L.J. and N.C. Partridge, Cellular and molecular mechanisms of bone

remodeling. Journal of Biological Chemistry, 2010. 285(33): p. 25103-25108.

17. Feng, X. and J.M. McDonald, Disorders of bone remodeling. Annual review of

pathology, 2011. 6: p. 121-145.

18. Peel, N., Disorders of bone metabolism. Surgery (Oxford), 2012. 30(2): p. 61-66.

19. Whyte, M.P., Paget's disease of bone. New England Journal of Medicine, 2006. 355(6):

p. 593-600.

20. Organization, W.H., http://www.who.int/inf-pr-1999/en/pr99-58.html. 2014,

consulted at 6/1/2014.

21. Allareddy, V., et al., Emergency Department Visits With Facial Fractures Among

Children and Adolescents: An Analysis of Profile and Predictors of Causes of Injuries.

Journal of Oral and Maxillofacial Surgery, 2014. 72(9): p. 1756-1765.

22. Zimmermann, G. and A. Moghaddam, Allograft bone matrix versus synthetic bone

graft substitutes. Injury, 2011. 42: p. S16-S21.

23. Bhatt, R.A. and T.D. Rozental, Bone graft substitutes. Hand clinics, 2012. 28(4): p.

457-468.

24. Andrews, D., G. Scholes, and G. Wiederrecht, Comprehensive Nanoscience and

Technology, Five-Volume set: Online Version. 2010: Academic Press.

25. Salgado, A.J., et al., Tissue Engineering and Regenerative Medicine: Past, Present,

and Future. Tissue engineering of the peripheral nerve: stem cells and regeneration

promoting factors, 2013. 108: p. 1-33.

26. Forgács, G. and W. Sun, Biofabrication: micro-and nano-fabrication, printing,

patterning and assemblies. 2013: William Andrew.

27. Verma, P. and V. Verma, Concepts of Tissue Engineering, in Animal biotechnology.

2014. p. 233-245.

28. Valente, J., et al., Alginate based scaffolds for bone tissue engineering. Materials

Science and Engineering: C, 2012. 32(8): p. 2596-2603.

Page 66: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

49

29. Lanza, R., R. Langer, and J.P. Vacanti, Principles of tissue engineering. 2011:

Academic press.

30. Hutmacher, D.W., et al., State of the art and future directions of scaffold‐based

bone engineering from a biomaterials perspective. Journal of tissue engineering and

regenerative medicine, 2007. 1(4): p. 245-260.

31. Holzwarth, J.M. and P.X. Ma, Biomimetic nanofibrous scaffolds for bone tissue

engineering. Biomaterials, 2011. 32(36): p. 9622-9629.

32. Holzapfel, B.M., et al., How smart do biomaterials need to be? A translational

science and clinical point of view. Advanced drug delivery reviews, 2013. 65(4): p.

581-603.

33. Albrektsson, T., et al., Osseointegrated titanium implants: requirements for ensuring

a long-lasting, direct bone-to-implant anchorage in man. Acta Orthopaedica, 1981.

52(2): p. 155-170.

34. Albrektsson, T. and M. Jacobsson, Bone-metal interface in osseointegration. The

Journal of Prosthetic Dentistry, 1987. 57(5): p. 597-607.

35. Stevens, B., et al., A review of materials, fabrication methods, and strategies used

to enhance bone regeneration in engineered bone tissues. Journal of Biomedical

Materials Research Part B: Applied Biomaterials, 2008. 85(2): p. 573-582.

36. Liu, Y., J. Lim, and S.-H. Teoh, Review: Development of clinically relevant scaffolds

for vascularised bone tissue engineering. Biotechnology advances, 2013. 31(5): p.

688-705.

37. Puppi, D., et al., Polymeric materials for bone and cartilage repair. Progress in

Polymer Science, 2010. 35(4): p. 403-440.

38. Liu, C., Z. Xia, and J. Czernuszka, Design and development of three-dimensional

scaffolds for tissue engineering. Chemical Engineering Research and Design, 2007.

85(7): p. 1051-1064.

39. Qiao, P., et al., Injectable calcium phosphate-alginate-chitosan microencapsulated

MC3T3-E1 cell paste for bone tissue engineering in vivo. Materials science &

engineering. C, Materials for biological applications, 2013. 33(8): p. 4633-4639.

40. Pawar, S.N. and K.J. Edgar, Alginate derivatization: a review of chemistry,

properties and applications. Biomaterials, 2012. 33(11): p. 3279-3305.

Page 67: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

50

41. Han, J., et al., Alginate–chitosan/hydroxyapatite polyelectrolyte complex porous

scaffolds: preparation and characterization. International journal of biological

macromolecules, 2010. 46(2): p. 199-205.

42. Lee, K.Y. and D.J. Mooney, Alginate: Properties and biomedical applications.

Progress in Polymer Science, 2012. 37(1): p. 106-126.

43. Morais, D., et al., Development and characterization of novel alginate-based

hydrogels as vehicles for bone substitutes. Carbohydrate polymers, 2013. 95(1): p.

134-142.

44. Christensena, B.E., Alginates as biomaterials in tissue engineering. Carbohydrate

Chemistry: Chemical and Biological Approaches, 2011. 37: p. 227-258.

45. Rodrigues, A.P., et al., The influence of preparation conditions on the characteristics

of chitosan‐alginate dressings for skin lesions. Journal of applied polymer science,

2008. 109(4): p. 2703-2710.

46. Kong, M., et al., Antimicrobial properties of chitosan and mode of action: a state of

the art review. International journal of food microbiology, 2010. 144(1): p. 51-63.

47. Di Martino, A., M. Sittinger, and M.V. Risbud, Chitosan: A versatile biopolymer for

orthopaedic tissue-engineering. Biomaterials, 2005. 26(30): p. 5983-5990.

48. Dash, M., et al., Chitosan—A versatile semi-synthetic polymer in biomedical

applications. Progress in Polymer Science, 2011. 36(8): p. 981-1014.

49. Miguel, S.P., et al., Thermoresponsive chitosan-agarose hydrogel for skin

regeneration. Carbohydrate Polymers, 2014. 111: p. 366-373.

50. Hutmacher, D., J. Goh, and S. Teoh, An introduction to biodegradable materials for

tissue engineering applications. Annals of the Academy of Medicine, Singapore, 2001.

30(2): p. 183-191.

51. Jiang, T., et al., Micro-and nanofabrication of chitosan structures for regenerative

engineering. Acta biomaterialia, 2014. 10(4): p. 1632-1645.

52. Croisier, F. and C. Jérôme, Chitosan-based biomaterials for tissue engineering.

European Polymer Journal, 2013. 49(4): p. 780-792.

53. Sarmento, B. and J. das Neves, Chitosan-based systems for biopharmaceuticals:

delivery, targeting and polymer therapeutics. 2012: John Wiley & Sons.

Page 68: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

51

54. Costa-Pinto, A.R., R.L. Reis, and N.M. Neves, Scaffolds based bone tissue

engineering: the role of chitosan. Tissue Engineering Part B: Reviews, 2011. 17(5): p.

331-347.

55. Li, Z., et al., Chitosan–alginate hybrid scaffolds for bone tissue engineering.

Biomaterials, 2005. 26(18): p. 3919-3928.

56. Jin, H.-H., et al., In-situ formation of the hydroxyapatite/chitosan-alginate

composite scaffolds. Materials Letters, 2008. 62(10–11): p. 1630-1633.

57. Jin, H.-H., et al., < i> In vivo</i> evaluation of porous hydroxyapatite/chitosan–

alginate composite scaffolds for bone tissue engineering. International journal of

biological macromolecules, 2012. 51(5): p. 1079-1085.

58. Diogo, G., et al., Manufacture of β-TCP/alginate scaffolds through a Fab@ home

model for application in bone tissue engineering. Biofabrication, 2014. 6(2): p.

025001.

59. Sowjanya, J., et al., Biocomposite scaffolds containing chitosan/alginate/nano-silica

for bone tissue engineering. Colloids and Surfaces B: Biointerfaces, 2013. 109: p. 294-

300.

60. Ryu, H.-S., et al., An improvement in sintering property of< i> β</i>-tricalcium

phosphate by addition of calcium pyrophosphate. Biomaterials, 2002. 23(3): p. 909-

914.

61. Lee, J.T.Y., et al., Comparative in vitro osteoinductivity study of CaP ceramics (HA,

α-TCP, β-TCP) using 10T1/2 cells with different controls and possible correlations

with other systems. Journal of Biomaterials and Nanobiotechnology, 2011. 2(02): p.

162.

62. Ma, P.X., Biomimetic materials for tissue engineering. Advanced drug delivery

reviews, 2008. 60(2): p. 184-198.

63. Torres, A., et al., Bioactive polymeric–ceramic hybrid 3D scaffold for application in

bone tissue regeneration. Materials Science and Engineering: C, 2013. 33(7): p. 4460-

4469.

64. Santos, C.F., et al., Design and production of sintered β-tricalcium phosphate 3D

scaffolds for bone tissue regeneration. Materials Science and Engineering: C, 2012.

32(5): p. 1293-1298.

Page 69: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

52

65. Weinand, C., et al., Hydrogel-β-TCP scaffolds and stem cells for tissue engineering

bone. Bone, 2006. 38(4): p. 555-563.

66. Chen, B., et al., Fabrication and mechanical properties of β-TCP pieces by gel-casting

method. Materials Science and Engineering: C, 2008. 28(7): p. 1052-1056.

67. Coelho, P.G., et al., Physico/chemical characterization and preliminary human

histology assessment of a β-TCP particulate material for bone augmentation.

Materials Science and Engineering: C, 2009. 29(7): p. 2085-2091.

68. Chang, H.-I. and Y. Wang, Cell responses to surface and architecture of tissue

engineering scaffolds. Regenerative Medicine and Tissue Engineering—Cells and

Biomaterials, 2011: p. 569-588.

69. Karageorgiou, V. and D. Kaplan, Porosity of 3D biomaterial scaffolds and

osteogenesis. Biomaterials, 2005. 26(27): p. 5474-5491.

70. Luo, Y., et al., 3D Scaffolds. Principles of tissue engineering, 2000: p. 475-494.

71. Bose, S., S. Vahabzadeh, and A. Bandyopadhyay, Bone tissue engineering using 3D

printing. Materials Today, 2013. 16(12): p. 496-504.

72. Wu, X., et al., Preparation of aligned porous gelatin scaffolds by unidirectional

freeze-drying method. Acta biomaterialia, 2010. 6(3): p. 1167-1177.

73. Fu, H., et al., In vitro evaluation of borate-based bioactive glass scaffolds prepared

by a polymer foam replication method. Materials Science and Engineering: C, 2009.

29(7): p. 2275-2281.

74. Muhamad Nor, M.A.A., et al., Preparation and characterization of ceramic foam

produced via polymeric foam replication method. Journal of materials processing

technology, 2008. 207(1): p. 235-239.

75. Fu, Q., et al., Mechanical and in vitro performance of 13–93 bioactive glass scaffolds

prepared by a polymer foam replication technique. Acta Biomaterialia, 2008. 4(6): p.

1854-1864.

76. Tampieri, A., et al., Porosity-graded hydroxyapatite ceramics to replace natural

bone. Biomaterials, 2001. 22(11): p. 1365-1370.

Page 70: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

53

77. Vivanco, J., et al., Mechanical characterization of injection-molded macro porous

bioceramic bone scaffolds. Journal of the mechanical behavior of biomedical

materials, 2012. 9: p. 137-152.

78. Sanosh, K., et al., Sol–gel synthesis of pure nano sized β-tricalcium phosphate

crystalline powders. Current Applied Physics, 2010. 10(1): p. 68-71.

79. Fukasawa, T., et al., Pore structure of porous ceramics synthesized from water-

based slurry by freeze-dry process. Journal of Materials Science, 2001. 36(10): p.

2523-2527.

80. Subia, B., J. Kundu, and S. Kundu, Biomaterial scaffold fabrication techniques for

potential tissue engineering applications. Tissue engineering, 2010.

81. Ho, M.-H., et al., Preparation of porous scaffolds by using freeze-extraction and

freeze-gelation methods. Biomaterials, 2004. 25(1): p. 129-138.

82. Mao, J.S., et al., Structure and properties of bilayer chitosan–gelatin scaffolds.

Biomaterials, 2003. 24(6): p. 1067-1074.

83. Seol, Y.-J., et al., Chitosan sponges as tissue engineering scaffolds for bone

formation. Biotechnology letters, 2004. 26(13): p. 1037-1041.

84. Li, Z. and M. Zhang, Chitosan–alginate as scaffolding material for cartilage tissue

engineering. Journal of Biomedical Materials Research Part A, 2005. 75(2): p. 485-

493.

85. Kang, K., L. Hockaday, and J. Butcher, Quantitative optimization of solid freeform

deposition of aqueous hydrogels. Biofabrication, 2013. 5(3): p. 035001.

86. Taqieddin, E. and M. Amiji, Enzyme immobilization in novel alginate–chitosan core-

shell microcapsules. Biomaterials, 2004. 25(10): p. 1937-1945.

87. Yeong, W.-Y., et al., Rapid prototyping in tissue engineering: challenges and

potential. Trends in biotechnology, 2004. 22(12): p. 643-652.

88. Cohen, D.L., et al., Increased mixing improves hydrogel homogeneity and quality of

three-dimensional printed constructs. Tissue Engineering Part C: Methods, 2010.

17(2): p. 239-248.

89. Lixandrão Filho, A., et al., Construction and adaptation of an open source rapid

prototyping machine for biomedical research purposes-a multinational collaborative

Page 71: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

54

development. Innovative developments in design and manufacturing. Leiria, Portugal:

CRC Press, 2009.

90. Lv, Q. and Q. Feng, Preparation of 3-D regenerated fibroin scaffolds with freeze

drying method and freeze drying/foaming technique. Journal of Materials Science:

Materials in Medicine, 2006. 17(12): p. 1349-1356.

91. Kang, H.-W., Y. Tabata, and Y. Ikada, Fabrication of porous gelatin scaffolds for

tissue engineering. Biomaterials, 1999. 20(14): p. 1339-1344.

92. Nie, H.-L. and L.-M. Zhu, Adsorption of papain with Cibacron Blue F3GA carrying

chitosan-coated nylon affinity membranes. International journal of biological

macromolecules, 2007. 40(3): p. 261-267.

93. Correia, T.R., et al., A bi-layer electrospun nanofiber membrane for plasmid DNA

recovery from fermentation broths. Separation and Purification Technology, 2013.

112: p. 20-25.

94. Coimbra, P., et al., Sodium hyaluronate/chitosan polyelectrolyte complex scaffolds

for dental pulp regeneration: synthesis and characterization. International journal of

biological macromolecules, 2011. 49(4): p. 573-579.

95. Lam, C.X., et al., Evaluation of polycaprolactone scaffold degradation for 6 months

in vitro and in vivo. Journal of biomedical materials research part A, 2009. 90(3): p.

906-919.

96. Lacroix, D., et al., Micro-finite element models of bone tissue-engineering scaffolds.

Biomaterials, 2006. 27(30): p. 5326-5334.

97. Noriega, S.E. and A. Subramanian, Consequences of Neutralization on the

proliferation and cytoskeletal organization of chondrocytes on Chitosan-

BasedMatrices. International Journal of Carbohydrate Chemistry, 2011: p. 13 pages.

98. Marques, J.G., et al., Synthesis and characterization of micelles as carriers of non-

steroidal anti-inflammatory drugs (NSAID) for application in breast cancer therapy.

Colloids and Surfaces B: Biointerfaces, 2014. 113: p. 375-383.

99. Ribeiro, M., et al., Dextran-based hydrogel containing chitosan microparticles loaded

with growth factors to be used in wound healing. Materials Science and Engineering:

C, 2013. 33(5): p. 2958-2966.

Page 72: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

55

100. Kim, K., et al., Incorporation and controlled release of a hydrophilic antibiotic using

poly (lactide-co-glycolide)-based electrospun nanofibrous scaffolds. Journal of

Controlled Release, 2004. 98(1): p. 47-56.

101. Shrivats, A.R., M.C. McDermott, and J.O. Hollinger, Bone tissue engineering: state of

the union. Drug discovery today, 2014.

102. Gurr, M. and R. Mulhaupt, Rapid prototyping. Polymers for Advanced Functional

Materials, 2012. 8: p. 77-99.

103. Trachtenberg, J.E., F.K. Kasper, and A.G. Mikos, Polymer scaffold fabrication.

Principles of Tissue Engineering, 2013: p. 423-440.

104. Villetti, M., et al., Thermal degradation of natural polymers. Journal of thermal

analysis and calorimetry, 2002. 67(2): p. 295-303.

105. Beyler, C.L. and M.M. Hirschler, Thermal decomposition of polymers. SFPE handbook

of fire protection engineering, 2002. 2: p. 110-131.

106. Phang, Y.-N., et al., Thermal and microbial degradation of alginate-based

superabsorbent polymer. Polymer Degradation and Stability, 2011. 96(9): p. 1653-

1661.

107. Douglas de Britto and Sergio Paulo Campana-Filho, Kinetics of the thermal

degradation of chitosan. Thermochimica acta, 2007. 465(1): p. 73-82.

108. Knaack, D., et al., Resorbable calcium phosphate bone substitute. Journal of

biomedical materials research, 1998. 43(4): p. 399-409.

109. Yang, F., et al., Poly (l, l-lactide-< i> co</i>-glycolide)/tricalcium phosphate

composite scaffold and its various changes during degradation in vitro. Polymer

degradation and stability, 2006. 91(12): p. 3065-3073.

110. Carano, R.A. and E.H. Filvaroff, Angiogenesis and bone repair. Drug discovery today,

2003. 8(21): p. 980-989.

111. Metsger, D., M. Rieger, and D. Foreman, Mechanical properties of sintered

hydroxyapatite and tricalcium phosphate ceramic. Journal of Materials Science:

Materials in Medicine, 1999. 10(1): p. 9-17.

112. Levy, P.-Y., et al., Relation entre portage nasal de< i> S</i>.< i> aureus</i> et

infections du site opératoire en orthopédie. Rôle de la contamination nasale: revue

Page 73: A influência do processo de produção de scaffolds na ... · diferentes: prototipagem rápida com uma impressora 3D Fab@home, liofilização e método de replicação de esponja.

56

systématique et méta-analyse de la littérature. Revue de Chirurgie Orthopédique et

Traumatologique, 2013. 99(6): p. 535-542.